Next Article in Journal
A Novel Non-Intrusive Vibration Energy Harvesting Method for Air Conditioning Compressor Unit
Next Article in Special Issue
Effects of Silicic Acid on Leaching Behavior of Arsenic from Spent Calcium-Based Adsorbents with Arsenite
Previous Article in Journal
Can the Market Deliver 100% Organic Seed and Varieties in Europe?
Previous Article in Special Issue
Current Management Condition and Waste Composition Characteristics of Construction and Demolition Waste Landfills in Hanoi of Vietnam
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Steel Slag and Autoclaved Aerated Concrete Grains as Low-Cost Adsorbents to Remove Cd2+ and Pb2+ in Wastewater: Effects of Mixing Proportions of Grains and Liquid-to-Solid Ratio

by
Gajanayake Mudalige Pradeep Kumara
1,* and
Ken Kawamoto
2,3
1
Department of Biosystems Technology, Faculty of Technology, University of Sri Jayewardenepura, Nugegoda 10250, Sri Lanka
2
Graduate School of Science and Engineering, Saitama University, 255 Shimo-Okubo, Sakura-ku, Saitama 3388570, Japan
3
Faculty of Building and Industrial Construction, Hanoi University of Civil Engineering, No. 55 Giai Phong Street, Hai Ba Trung District, Hanoi 11616, Vietnam
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(18), 10321; https://doi.org/10.3390/su131810321
Submission received: 17 August 2021 / Revised: 13 September 2021 / Accepted: 13 September 2021 / Published: 15 September 2021
(This article belongs to the Special Issue Environmentally Sound Waste Management and Zero Waste Principles)

Abstract

:
This study investigated the applicability of industrial by-products such as steel slag (SS) and autoclaved aerated concrete (AAC) grains (<0.105, 0.105–2, 2–4.75 mm) as low-cost adsorbents for simultaneous removal of Cd2+ and Pb2+ in wastewater. A series of batch adsorption experiments was carried out in single and binary-metal solutions of Cd2+ and Pb2+ by changing the mixing proportions of SS and AAC grains. In addition, the effect of the liquid-to-solid ratio (L/S) on the removal of Cd2+ and Pb2+ in multi-metal solution was examined. Results showed that SS grains had a high affinity with Cd2+ in the single solution, while AAC grains had an affinity with Pb2+. In the binary solution, the mixtures of SS and AAC grains removed both Cd2+ and Pb2+ well; especially, the tested adsorbents of SS+AAC [1:1] and SS+AAC [1:4] mixtures achieved approximately 100% removal of both metals. Based on the results in the multi-metal solutions, the metal removal % and selectivity sequence varied depending on the mixed proportions of SS and AAC grains and L/S values. It was found that the SS+AAC [1:1] mixture of SS and AAC grains showed 100% removals of Cd2+, Pb2+, Cu2+, Ni2+, and Zn2+ simultaneously at L/S = 10 and 60.

1. Introduction

Due to rapid urbanization and industrialization, the discharge of wastewater containing toxic heavy metals from various sources such as industries, mines, vehicles, batteries, and metal-containing paints is increasing all over the world, especially in developing countries [1,2]. Heavy metal ions are non-biodegradable and tend to accumulate in living organisms, and those are considered toxic or carcinogenic ions [3,4]. Over the past five decades, for example, the annual global release of heavy metals reached 22,000 tons for cadmium, 939,000 tons for copper, 783,000 tons for lead, and 1,350,000 tons for zinc [5]. Inadequate treatments of toxic heavy metals such as Cd2+ and Pb2+ cause not only serious surface and groundwater pollutions and soil contamination [6,7,8] but also various human diseases including acute or chronic poisoning, dermatitis, brain damage in children, and digestive tract cancer [9,10,11]. Furthermore, the direct discharge of Cd2+ and Pb2+ into the sewage system causes a negative impact on the effectiveness of biological wastewater treatment [12]. Thus, Cd2+ and Pb2+ can be considered the most commonly available and most harmful heavy metals to humans as well as the environment. Thus, proper treatment of these heavy metals in wastewater is essential to protect public health and the environment.
Numerous efforts are being made to develop cost-effective innovative methods of wastewater treatment, such as hydroxide precipitation [13], membrane filtration [14], reverse osmosis [15], phytoextraction [16], ion exchange [17], electrokinetic remediation [18], and adsorption [19]. When developing new methods, economic feasibility and sustainability are given high priority. However, adsorption has received a great deal of attention because of its high efficiency and cost effectiveness [20]. At present, a number of adsorbents are used for heavy metal removal on experimental as well as industrial levels. Out of those, activated carbon is the most widely used adsorbent due to its high adsorption capacity [21]. However, activated carbon is still in limited use because of economic reasons, mostly in developing countries [22]. Therefore, the evaluation of the effective use of low-cost and locally available adsorbents is receiving much attention. Low-cost adsorbents are classified generally as geo-sorbents [23,24,25,26,27], bio-sorbents [28,29,30,31,32], industrial wastes [33,34,35], industrial by-products (IBPs) and construction and demolition waste (CDW) [36,37,38,39,40,41,42], and modified low-cost adsorbents [19,21,43,44,45,46]. Many adsorption experiments have been performed to examine their adsorption capacities and characteristics [47]. However, among these low-cost adsorbents, IBPs and CDW are still marginalized in adsorption studies as well as industrial applications, although those materials have the same potential as other low-cost adsorbents to adsorb heavy metals from wastewater [36,37,38,39,48,49,50]. Therefore, more studies on the effective use of IBPs and CDW are required.
The growing industrialization in developing countries, on the other hand, generates large amounts of IBPs such as steel slag (SS) by the steel-making industry, coal ash by coal-fired power plants, and mining waste by extracting and processing mineral resources [51,52]. Along with the rapid generation of IBPs, improper treatment and uncontrolled dumping of IBPs are becoming a major environmental issue in many developing countries [53,54]. Moreover, with increasing production of non-fired bricks such as cement blocks and autoclaved aerated concrete (AAC) all over the world, the amount of scrap waste has also been increasing in developing countries [55,56]. The AAC scrap waste is not fully reused in most of the countries but dumped on-site without any treatment. Proper management is vital for sustainable development, and effective utilizations of IBPs and scrap waste of non-fired bricks are required for circulating those materials in society. Hence, it is timely and important to promote the reuse of these abundantly available waste materials as a cost-effective adsorbent for wastewater treatment process to give added value to IBPs and AAC scrap waste.
Until now, many studies have investigated the applicability of IBPs and other types of low-cost adsorbents for heavy metal treatment processes, but most of the studies mainly target the treatment of heavy metals at low concentrations (typically, heavy metal concentrations <200 mg/L) [21,57,58,59]. Furthermore, most of these studies examined the adsorption capacities of target adsorbents in single and/or binary-metal solutions at a constant liquid-to-solid (L/S) ratio [50]. Because actual industrial wastewater might be contaminated by multi-species heavy metals like Pb2+, Cd2+, Ni2+, Cu2+, and Zn2+, further studies are needed to examine the simultaneous removal of heavy metals fully considering the co-existence of competitive metals in wastewater [19,60]. For example, it has been reported that Cd2+ adsorption onto cementitious adsorbents, including AAC grains, was hampered by the co-existence of Pb2+ and Cu2+ due to their high affinities with the tested adsorbents [61,62,63,64]. In addition, Kumara et al. 2019c [65] examined the simultaneous removal potential of Cd2+ and Pb2+ in the multi-metal solutions by AAC, SS, and AAC+SS [1:1] grains and revealed selectivity sequences of Pb2+ > Cu2+ > Ni2+ > Zn2+ > Cd2+ for AAC, Pb2+ > Cu2+ > Cd2+ > Ni2+ > Zn2+ for SS, and Pb2+ ≈ Cu2+ > Cd2+ > Zn2+ > Ni2+ for AAC+SS [1:1] under an L/S ratio 60. Thus, many researchers have concluded that the simultaneous removal of Cd2+-like metals using a low-cost adsorbent is impossible. Therefore, it is worthwhile to discover low-cost adsorbents that have potential for simultaneous removal of commonly available heavy metals in wastewater. In order to examine the simultaneous removal of Cd2+ and Pb2+ in wastewater, therefore, this study investigated the adsorption characteristics of those metals onto the mixtures of SS and AAC grains in single and binary-metal solutions. Especially, the selectivity sequence on metal adsorption was carefully investigated in the multi-metal solutions (coexistence of Ni2+, Cu2+, and Zn2+) at different L/S ratios ranging from 5 to 250.

2. Materials and Methods

2.1. Adsorbents Preparation and Characterization

Commercially available steel slag (SS) for civil engineering applications (Nippon Steel Cooperation and Sumitomo Metal Industries, Ltd., Saitama, Japan) and autoclaved aerated concrete (AAC) (Asahi Kasei Construction Material Corp., Tokyo, Japan) were used. General information on the manufacturing and characteristics of SS and AAC are given in Nippon Slag Association [66] and Trong et al. [67]. After crushing by hand in the laboratory, the tested samples were sieved to three grain fractions of <0.105, 0.105–2, and 2–4.75 mm. The dataset of material properties and the adsorption tests for AAC grains analyzed in this study were obtained from Kumara et al., 2019a [61].
The basic physical and chemical properties of SS and AAC grains are summarized in Table 1. The Brunauer–Emmett–Teller (BET) surface area (SSA) was measured by a ASAP2020 adsorption analyzer (Micromeritics, Norcross, GA, USA). Measured specific gravity (Gs) values of SS grains were higher compared to AAC grains, and pH values showed that SS grains are alkaline in water due to the hydration reaction of CaO with a release of OH-. The SSA values of SS grains were lower than those of AAC grains and decreased with decreasing grain size. This implied that SS grains had fewer inter-grain pores compared to AAC grains, and the outer-grain surface controlled the SSA of SS grains.
The chemical composition of tested adsorbents was characterized by energy-dispersive X-ray spectroscopy (EDX; X-Max Extreme, Oxford Instruments, High Wycombe, UK) and X-ray diffractometry (XRD; XRD-7000, Shimadzu Cooperation, Kyoto, Japan). The EDX test data are shown in Table 2. Higher CaO (41.7%) and Fe2O3 (22.5%), and lower SiO2 (16.8%) were found for SS grains compared to those of AAC grains, indicating a potential for the ion exchange (Ca2+) reaction with heavy metals in wastewater like other calcium silicate materials [68,69]. The XRD analysis showed that magnetite (Fe3O4), iron oxidate (FeO), calcite (CaCO3), calcium hydroxide (Ca(OH)2), and halloysite (Al2Si2O5(OH)4) were the main minerals in SS grains. Both EDX and XRD analyses confirmed the existence of high amounts of metal oxides and hydroxides in SS grains, implying that SS grains favored the adsorption of heavy metals such as Cd2+, Cu2+, and Pb2+ [70,71,72].

2.2. Procedures of Batch Experiments

Stock solutions of Cd2+, Pb2+, Cu2+, Ni2+, and Zn2+ (synthesized wastewater) were prepared by dissolving CdCl2, PbCl2, CuCl2, NiCl2, and ZnCl2, respectively, in deionized water (DI). The solutions’ pH and ionic strength were controlled using 1N HCl, 1N NaOH, and NaNO3. All chemicals used for this study were analytical grade from Wako Pure Chemical Industries, Ltd., Osaka, Japan, with more than 98% chemical purity. A standard batch method recommended by the Organization of Economic Cooperation and Development [73] was used for all batch adsorption experiments. The concentrations of heavy metals and other ions were measured using a flame atomic absorption spectrophotometer (AAS; AA 6200, Shimadzu, Japan), the exact pH and EC values for each metal solution were measured before the experiment, and their changes were observed during the experiments using a pH/EC meter.
Test conditions of adsorption experiments are summarized in Table 3. The types of batch adsorption experiments in this study were categorized into four groups to investigate: (1) adsorption isotherm in a single-metal solution with L/S = 60 for Cd2+ and 10 for Pb2+, (2) % removal in a binary-metal solution (Cd2+, Pb2+) with L/S = 60, (3) % removal in multi-metal solution (Cd2+, Pb2+, Cu2+, Ni2+, Zn2+) with L/S = 60, and (4) the effect of L/S ratios on % removal in a multi-metals solution with different L/S ratios = 5–250.
Three grain sizes of adsorbents of <0.105, 0.105–2, and 2–4.75 mm were used for the adsorption isotherm experiments in a single-metal solution, and an adsorbent grain size of 0.105–2 mm was used for other adsorption experiments. All test conditions were carried out with triplicate measurements, and the averaged values were given in the paper because of a small variation of measured data.

2.2.1. Adsorption Isotherms for Cd2+ and Pb2+ in a Single-Metal Solution

Adsorption isotherm experiments for Cd2+ and Pb2+ onto SS and AAC grains were carried out at natural pH with the initial metal concentration (Ci) of 0–5000 mg/L for Cd2+ and 0–1500 mg/L for Pb2+ (19 different concentrations, fully considering the actual metal concentration range in industrial wastewater [74]) at L/S = 60. To determine the maximum adsorption capacity and intensity of Cd2+ and Pb2+ onto the tested adsorbents, Langmuir (Equation (1)) and Freundlich (Equation (2)) models were used to fit the obtained experimental data:
Ce/Qe = 1/(bQm)+ Ce/Qm
log Qe = log Kf + 1/n (log Ce)
where Ce (mg/L) is the equilibrium concentration of heavy metals, Qe (mg/g) is the amount adsorbed per adsorbent at the equilibrium, b (g/L) is the Langmuir constant related to binding strength, Qm (mg/g) is the maximum adsorption capacity, Kf (L/g) is the Freundlich adsorption capacity, and 1/n is the adsorption intensity.

2.2.2. Effect of Competitive Metal Ions on Cd2+ and Pb2+ Adsorption in Binary Metal Solution

To examine the effect of competitive metal ions on Cd2+ and Pb2+ adsorption, batch experiments were carried out in a binary-metal solution of Cd2+ and Pb2+ at natural pH with Ci = 1000 mg/L. Five different mixtures of SS and AAC grains, i.e., mixing proportions of SS (alone), SS+AAC [4:1], SS+AAC [1:1], SS+AAC [1:4], and AAC (alone), were used as tested adsorbents. The ratios in brackets show the mixing proportions in weight % (e.g., (1:1) means a mixture of 50% SS and 50% AAC). In binary-metal adsorption experiments, the mixed solutions of Cd2+ to Pb2+ were used at metal molar ratios of 1:0, 1:0.25, 1:0.5, 1:0.75, 1:1, 1:2, and 1:5 to investigate the effect of Pb2+ concentrations on Cd2+ adsorption. Vice versa, the same mixing solutions of Pb2+ and Cd2+ solutions were used to investigate the effect of Cd2+ on Pb2+ adsorption. In each experiment, the metal removal percentage (R, %) was calculated using Equation (3) [46]:
R = [(Ci–Ce)/Ci] × 100

2.2.3. Effect of Competitive Metal Ions on Cd2+ and Pb2+ Adsorption in Multi-Metal Solutions

To examine the effects of competitive metals on Cd2+ and Pb2+ adsorption onto tested adsorbents with different L/S, batch adsorption experiments in multi-metal solutions were carried out using a mixed metal solution of Cd2+, Pb2+, Cu2+, Ni2+, and Zn2+ at natural pH with Ci = 1000 mg/L. Like the adsorption experiments of binary-metal solutions, five different mixtures of SS and AAC grains (SS, SS+AAC [4:1], SS+AAC [1:1], SS+AAC [1:4], and AAC) were used with five different L/S values of 5, 10, 60, 100, and 250. In each experiment, the R value of each metal was calculated using Equation (3).

3. Results and Discussion

3.1. Adsorption Isotherms for Cd2+ and Pb2+ in Single Metal Solution

The measured adsorption isotherms for Cd2+ and Pb2+ onto SS grains with different grain sizes are shown in Figure 1. In the figures, the Langmuir model (Equation (1)) was well-fitted to the measured data except for the Cd2+ adsorption onto SS grains with <0.105 mm (Figure 1a). The SS grain with <0.105 mm showed a very high Cd2+ adsorption and did not show the maximum adsorption capacity (Qm) in the range of Ci = 0–5000 mg/L. For both Cd2+ and Pb2+ adsorptions, the adsorption decreased with increasing grain size. This can be understood to indicate that the adsorption surface area of tested adsorbents controlled the adsorption capacity, and the sample of <0.105 mm with higher SSA imposed a higher adsorption capacity compared to the samples with lower SSA of the grain sizes of 0.105–2, 2–4.75 mm (Table 1). The Freundlich model (Equation (2)), on the other hand, fitted well with all tested adsorbents in the range of Ci = 0–5000 mg/L for Cd2+ and 0–1500 mg/L for Pb2+. This suggests that a monolayer adsorption is predominant at the early stage of the adsorption process (typically, Ci < 1500 mg/L) and the adsorption process shifts to a multilayer adsorption, especially at higher than Ci > 1500 mg/L, for both Cd2+ and Pb2+ [75,76].
The fitted adsorption parameters by Freundlich and Langmuir models for tested SS grains are summarized in Table 4 with the reported data for AAC grains [61]. It shows clearly that the SS grains have higher affinities (higher Qm and Kf) to Cd2+ compared to AAC grains while the AAC gains have higher affinities to Pb2+ compared to SS grains. These results imply that the mixing of SS and AAC grains would be effective for the simultaneous adsorption of Cd2+ and Pb2+, as discussed in the following sections. For reference, the measured Qm values in this study were compared to the previously reported values for different types of adsorbents such as IBPs (including construction and demolition waste) and geo- and bio-sorbents and are summarized in Table 5. It is noticeably shown that SS has the highest Qm of Cd2+ adsorption compared to those of other adsorbents.
Figure 2 shows the relationship between the amounts of Ca2+ and adsorbed metal released onto SS grains sized 0.105–2 mm in the measured Ci range. For both Cd2+ and Pb2+ adsorptions, the Ca2+ was released linearly along with the metal adsorption (r2 ≥ 0.88). The correlation regression showed that the ratios of released Ca2+ became greater than 1 (2.24 for Cd2+ and 6.37 for Pb2+). This means that approximately 2 and 6 times of Ca2+ were released when the one metal was adsorbed, indicating the adsorption mechanism was not controlled by a simple 1:1 ion exchange process of Ca2+ and Cd2+/Pb2+ on the adsorbent surface of SS grains in the single-metal solution system. On the other hand, Kumara et al. 2019a [61] observed an almost 1:1 relationship between the released Ca2+ amount and adsorbed Cd2+/Pb2+ for AAC grains, and the hydrated adsorbent surface was the dominant metal adsorption mechanism [77,78].
Table 5. Comparison of the maximum adsorption capacity (Qm) of tested adsorbents with reported values.
Table 5. Comparison of the maximum adsorption capacity (Qm) of tested adsorbents with reported values.
CategoryAdsorbentParticle Size (mm)BET Surface Area (m2/g)Liquid-to-Solid Ratio (L/S)pH RangeCd2+Pb2+Ref.
Ci Range (mg/L)Qm (mg/g)Ci Range (mg/L)Qm (mg/g)
Industrial by-products (IBPs) (including construction and demolition waste)Steel slag (SS)0.105–24.96010–120–5000130–3130–15008.2–17.5This study
Fly ash0.01–0.022.850–10,0002–100–53.80–55.1[35]
Blast furnace slag0.15–0.31.150–10,0002–100–55.10–54.9
Arc furnace slag0.9–23.421002–822.5–3606.541–66225[33]
Autoclaved aerated concrete 0.105–223.6608–1025–200016.525–2000257[61]
Crushed concrete fines0.105–211.2608–1125–200022.825–2000129
Crushed clay bricks0.105–215.9106–70–20003.20–20005.5[79]
Municipal solid waste slag0.105–2-108–90–20002.30–20003.3
Cellular concrete1.5–2.517714286.5–7.210–20028.710–200157[77]
Geo-sorbentsZeolite-15.440–10006.5–6.70–506.720–509.97[43]
Ca-bentonite<0.063871000≈51–10031.35–15085.5[80]
Sepiolite<0.1-100-50–60019.250–60050[81]
LimestonePowder11966672–61–20052.91–200184[82]
Montmorillonite--100≈210–5001.210–5003.25[83]
Bio-sorbentsCoconut coir husk<0.351.83400≈72.5–20047.325–20066.1[30]
Coconut husk/shell 0.07521210≈8.8100–20003.5100–200013.4[31]
Biochar0.1551.25005–750–9007550–900129[84]
Green algae1–1.5-1252–622–38239.241–70474.3[85]
Lemon berry0.15-1004.8–5110–2800129200–5200310[86]

3.2. Removal of Cd2+ and Pb2+ in Binary Metal Solution

Measured values of the percentage of metal removed, R % (Equation (3)) for Cd2+ and Pb2+ in the binary-metal solution by tested adsorbents, were plotted against the Cd:Pb and Pb:Cd molar mixed ratios in Figure 3. Like the test results in a single-metal solution for the removal of Cd2+, SS had a strong affinity with Cd2+ and became independent of the existence of Pb2+ in binary-metal solution, SS, and its mixtures (SS, SS+AAC [4:1], SS+AAC [1:1], SS+AAC [1:4]) gave approximately R = 100% (Figure 3a–d,f–i). In contrast, the removal of Pb2+ in the binary-metal solution was controlled by AAC, and AAC and its mixtures (AAC, SS+AAC [1:4], and SS+AAC [1:1]), which gave approximately R = 100% (Figure 3c–e,h–j). Previous studies reported that geo- and bio-adsorbents had difficulty achieving the sufficient removal of Cd2+ in the presence of Pb2+ in the binary-metal solution [30,31]. In this study, however, the tested results suggested strongly that the mixing of SS and AAC grains was effective to simultaneously remove Cd2+ and Pb2+ in a binary-metal solution; the mixtures of SS+AAC [1:1] and SS+AAC [1:4] were especially able to absorb Cd2+(or Pb2+) completely, even in the presence of Pb2+ (or Cd2+) in the solution.
In the batch adsorption tests in the binary-metal solution, the measured pH values after the adsorption (pHe) ranged from 8–10 for AAC grains (Figure 3e,j) and 10–12 for SS and mixed adsorbents of SS and AAC (Figure 3a–d,f–i). This indicated that the metals were adsorbed onto the tested adsorbents under alkaline conditions. Previous studies suggested that the metal adsorption onto cementitious and adsorbents rich in calcium metal oxides/hydroxides resulted in combined chemical processes and reactions under alkaline conditions: i.e., hydration of the adsorbent surface, hydrolysis of metal ions, physisorption, chemisorption, ion exchange, and surface complexation and precipitation [33,61,77,87,88,89,90]. For example, According to the metal speciation in pH-Eh diagram [91], Pb is presented as Pb2+ or possible to precipitate as Pb(OH)2 in the range of pH = 7–12 and most probably exists as Pb(OH)3- in the range of pH >12. On the other hand, Cd exists mainly in the forms of Cd2+ and Cd(OH)+ in the range of pH = 9–13. As shown in Table 2, the SS and AAC grains tested in this study were rich in CaO and Fe2O3. Based on the results of previous studies and the chemical compositions of tested adsorbents, the simultaneous removal of Cd2+ and Pb2+ from the mixtures of SS and AAC grains in the binary-metal solution under alkaline conditions can be understood as shown in Figure 4. For SS grains, under the given experimental conditions, active hydroxides (-OH) on the surface of metal oxides/hydroxides (especially Fe2O3) create a negative surface charge, resulting in the formation of an inner-sphere complex (surface complexation) by replacing Ca2+ (high affinity with Cd2+). For AAC grains, on the other hand, the ion exchange between Ca2+ and Pb2+ on the adsorbent surface becomes dominant, and the inner-sphere complex promotes the Pb2+ adsorption onto the surface with a negative charge (high affinity with Pb2+). Because these two reactions promote the release of Ca2+ ions, the amount of Ca2+ released became higher than the amount of metal adsorbed, as shown in Figure 2. In addition, hydrated CaO releases high amounts of OH- ions for both SS and AAC grains and promotes the precipitation of metal hydroxides (Cd(OH)2 and Pb(OH)2) under the alkaline condition.

3.3. Removal of Cd2+ and Pb2+ in Multi-Metal Solutions and Selectivity Sequence

Simultaneous removals of Cd2+ and Pb2+ in multi-metal solutions by the tested adsorbents were examined under different L/S ratios of 5, 10, 60, 100, and 250. The measured R values of metal ions in multi-metal solutions for SS, SS+AAC [1:1], and AAC are exemplified in Figure 5, and selectivity sequences are shown in Table 6. For all tested adsorbents, the R values of metals became high in low L/S conditions and decreased with increasing L/S. Especially, the R values became low at L/S = 60, 100, and 250, except for the Pb2+ removal and the Cu2+ removal of SS and AAC grains at L/S = 60 (Figure 5). This suggests that control of L/S is an important factor to achieve high R values in multi-metal solutions, unlike the single-metal and binary-metal solutions.
As shown in Table 6, the selectivity sequences of metals in multi-metal solutions are highly dependent on the L/S for SS and AAC grains, and the Cd2+ and Pb2+ adsorptions are affected by the existence of competitive metal ions. The selectivity sequences for the mixtures of SS and AAC, on the other hand, were less dependent on the L/S compared to the results from single SS and AAC grains. For SS grains, the adsorption of Cd2+ (and Cu2+, Ni2+, and Zn2+) became much higher than that of Pb2+ at low L/S (=5 and 10), and the Pb2+ adsorption became higher than that of Cd2+. For AAC grains, the adsorption of Pb2+ (and Cu2+) became higher than that of Cd2+ at all L/S conditions in accordance with the tested results from single-metal and binary-metal solutions. For the mixtures of SS and AAC (SS+AAC [1:1]), the adsorption of Cd2+ became equal to the Pb2+ adsorption at low L/S (=5 and 10), but the sequence became the opposite at high L/S (=60, 100, and 250). Based on the test results above, therefore, the mixtures of SS and AAC grains are able to simultaneously remove Cd2+ and Pb2+ in multi-metals solution with high R values of approximately 100% at low L/S conditions (5 and 10). Again, the test results suggest the control of L/S is a key factor for the practical application of the mixtures of SS and AAC grains as adsorbents to remove Cd2+ and Pb2+ in wastewater.
In this study, SS exhibited a high capacity to adsorb Cd2+ from wastewater. For example, 0.105–2 mm particles of SS achieved effluent discharge standards of <0.001 mg/L for Cd2+ up to Ci < 1500 mg/L by the single-batch adsorption under the experimental conditions. Additionally, the SS+AAC (1:1) mixture removed 100% of Cd2+ and Pb2+ from binary and multi-metal solutions. However, the following facts should be considered in future studies and applications for the sustainable use of these low-cost adsorbents. Along with the metal adsorption process, SS and AAC grains released a relatively high concentration of Ca2+. Additionally, SS grains showed high alkalinity in water compared to AAC. Therefore, those factors should be controlled before treated water is discharged into the natural environment.

4. Conclusions

The results of batch adsorption experiments of the tested adsorbents revealed that SS grains had a high affinity for Cd2+ (>300 mg/g) but less affinity for Pb2+ (<20 mg/g). In the binary-metal solution, the mixtures of SS and AAC grains, especially SS+AAC [1:1] and SS+AAC [1:4], removed 100% of Cd2+ and Pb2+ simultaneously without any effect of metal molar mixed ratios at the L/S ratio of 60. Remarkably, in the multi-metal solutions, the metal R and selectivity sequence varied depending on the mixing proportions of SS and AAC grains and L/S values. It was found that the SS+AAC [4:1] and SS+AAC [1:1] mixtures of SS and AAC grains removed 100% of Cd2+, Pb2+, Cu2+, Ni2+, and Zn2+ simultaneously, at L/S = 10 and 60. The correlation regression showed that the ratios of released Ca2+ became greater than 1, meaning that approximately 2 and 6 times of Ca2+ were released in the one-metal adsorption process. For SS grains, active hydroxides (-OH) on the surface of metal oxides/hydroxides created a negative surface charge, resulting in the formation of an inner-sphere complex by replacing Ca2+. For AAC grains, the ion exchange between Ca2+ and Pb2+ on the adsorbent surface becomes dominant and the inner-sphere complex promotes the Pb2+ adsorption onto the surface with a negative charge. Thus, the amount of Ca2+ released became higher than the amount of metal adsorbed. Further studies are needed to examine the effects of other factors such as initial pH, temperature, and other competitive ions that control the heavy metal adsorption of SS and AAC grains. Additionally, the regeneration of metal-adsorbed adsorbents (i.e., collection of adsorption heavy metals from adsorbents) should be examined for practical application. The application of IBPs for the wastewater treatment, however, is essential with a high potential from the viewpoints of material efficiency and saving natural resource consumption.

Author Contributions

Conceptualization, K.K.; methodology, G.M.P.K.; software, G.M.P.K.; validation, K.K.; formal analysis, G.M.P.K.; investigation, K.K.; resources, K.K.; data curation, G.M.P.K.; writing—original draft preparation, G.M.P.K.; writing—review and editing, K.K.; visualization, G.M.P.K.; supervision, K.K.; project administration, K.K.; funding acquisition, K.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JST-JICA Science and Technology Research Partnership for Sustainable Development (SATREPS) Project (No. JPMJSA1701).

Institutional Review Board Statement

Not available.

Informed Consent Statement

Not available.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AACautoclaved aerated concrete
bLangmuir constant (g/L)
BETBrunauer–Emmett–Teller
CDWconstruction and demolition waste
Ceequilibrium concentration (mg/L)
Ciinitial concentration (mg/L)
DIdeionized water
ECelectrical conductivity (mS/cm)
EDXenergy-dispersive X-ray spectroscopy
Gsspecific gravity (-)
IBPsindustrial by-products
KfFreundlich adsorption capacity (L/g)
LOIloss on ignition (%)
L/Sliquid-to-solid ratio
nadsorption intensity (-)
pHeequilibrium pH
Qeequilibrium adsorption (mg/g)
Qmmaximum adsorption capacity (mg/g)
Rremoval percentage (%)
SSsteel slag
SSAsurface area (m2/g)
wADgravimetric water content at air-dry [g/g, in %]
XRDX-ray diffractometry

References

  1. Petrella, A.; Spasiano, D.; Acquafredda, P.; De Vietro, N.; Ranieri, E.; Cosma, P.; Rizzi, V.; Petruzzelli, V.; Petruzzelli, D. Heavy metals retention (Pb(II), Cd(II), Ni(II)) from single and multi-metal solutions by natural bio-sorbents from the olive oil milling operations. Process Saf. Environ. Prot. 2018, 114, 79–90. [Google Scholar] [CrossRef]
  2. Rad, L.R.; Momeni, A.; Ghazani, B.F.; Irani, M.; Mahmoudi, M.; Noghreh, B. Removal of Ni2+ and Cd2+ ions from aqueous solutions using electrospun PVA/zeolite nano-fibrous adsorbent. Chem. Eng. J. 2014, 256, 119–127. [Google Scholar] [CrossRef]
  3. Sellaoui, L.; Dotto, G.L.; Lamine, A.B.; Erto, A. Interpretation of single and competitive adsorption of cadmium and zinc on activated carbon using monolayer and exclusive extended monolayer models. Environ. Sci. Pollut. Res. 2017, 24, 19902–19908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Demey, H.; Vincent, T.; Guibal, E. A novel algal-based sorbent for heavy metal removal. Chem. Eng. J. 2018, 332, 582–595. [Google Scholar] [CrossRef]
  5. Singh, O.V.; Labana, S.; Pandey, G.; Budhiraja, R.; Jain, R.K. Phytoremediation: An overview of metallic ion decontamination from soil. Appl. Microb. Biotech. 2013, 61, 405–412. [Google Scholar] [CrossRef]
  6. Duan, Q.; Lee, J.; Liu, Y.; Chen, H.; Hu, H. Distribution of heavy metal pollution in surface soil samples in China: A graphical review. Bull. Environ. Contam. Toxicol. 2016, 97, 303–309. [Google Scholar]
  7. Wu, G.; Kang, H.; Zhang, X.; Shao, H.; Chu, L.; Ruan, C. A critical review on the bio-removal of hazardous heavy metals from contaminated soils: Issues, progress, eco-environmental concerns and opportunities. J. Hazard. Mater. 2010, 174, 1–8. [Google Scholar] [CrossRef] [PubMed]
  8. Elouear, Z.; Bouzid, J.; Boujelben, N. Removal of nickel and cadmium from aqueous solutions by sewage sludge ash: Study in single and binary systems. Environ. Technol. 2009, 30, 561–570. [Google Scholar] [CrossRef] [Green Version]
  9. Bertrand, P.S.J.; Silvestre, V.; Pinell, V. Lead-induced DNA damage in Vicia faba root cells: Potential involvement of oxidative stress. Mutat. Res./Genet. Toxicol. Environ. Mutagenesis 2011, 726, 123–128. [Google Scholar]
  10. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef] [PubMed]
  11. Khan, S.; Cao, Q.; Zheng, Y.M.; Huang, Y.Z.; Zhu, Y.G. Health risks of heavy metals in contaminated soils and food crops irrigated with wastewater in Beijing, China. Environ. Pollut. 2008, 152, 686–692. [Google Scholar] [CrossRef]
  12. Hui, S.; Chao, H.; Kot, C. Removal of mixed heavy metal ions in wastewater by zeolite 4A and residual products from recycled coal fly ash. J. Hazard. Mater. 2005, 127, 89–101. [Google Scholar] [CrossRef]
  13. Huisman, J.L.; Schouten, G.; Schultz, C. Biologically produced sulphide for purification of process streams, effluent treatment and recovery of metals in the metal and mining industry. Hydrome 2006, 83, 106–113. [Google Scholar] [CrossRef]
  14. Blöcher, C.; Dorda, J.; Mavrov, V.; Chmiel, H.; Lazaridis, N.K.; Matis, K.A. Hybrid flotation membrane filtration process for the removal of heavy metal ions from wastewater. Water Res. 2003, 37, 4018–4026. [Google Scholar] [CrossRef]
  15. Chan, B.K.C.; Dudeney, A.W.L. Reverse osmosis removal of arsenic residues from bioleaching of refractory gold concentrates. Miner. Eng. 2008, 21, 272–278. [Google Scholar] [CrossRef]
  16. Chehregani, A.; Noori, M.; Yazdi, H.L. Phytoremediation of heavy-metal-polluted soils: Screening for new accumulator plants in Angouran mine (Iran) and evaluation of removal ability. Ecotoxicol. Environ. Saf. 2009, 72, 1349–1353. [Google Scholar] [CrossRef]
  17. Alyüz, B.; Veli, S. Kinetics and equilibrium studies for the removal of nickel and zinc from aqueous solutions by ion exchange resins. J. Hazard. Mater. 2009, 167, 482–488. [Google Scholar] [CrossRef]
  18. Ouhadi, V.R.; Yong, R.N.; Shariatmadari, N.; Saeidijam, S.; Goodarzi, A.R.; Zanjani, M.S. Impact of carbonate on the efficiency of heavy metal removal from kaolinite soil by the electrokinetic soil remediation method. J. Hazard. Mater. 2010, 173, 87–94. [Google Scholar] [CrossRef]
  19. Li, Z.; Wang, L.; Meng, J.; Liu, X.; Xu, J.; Wang, F.; Brookes, P. Zeolite-supported nano-scale zero-valent iron: New findings on simultaneous adsorption of Cd(II), Pb(II), and As(III) in aqueous solution and soil. J. Hazard. Mater. 2018, 344, 1–11. [Google Scholar] [CrossRef] [PubMed]
  20. Mishra, P.C.; Patel, R.K. Removal of lead and zinc ions from water by low cost adsorbents. J. Hazard. Mater. 2009, 168, 319–325. [Google Scholar] [CrossRef]
  21. Karnib, M.; Kabbanib, A.; Holail, H.; Olama, Z. Heavy metals removal using activated carbon, silica and silica activated carbon composite. Energy Procedia 2014, 50, 113–120. [Google Scholar] [CrossRef] [Green Version]
  22. Bailey, S.E.; Olin, T.J.; Bricka, R.M.; Adrian, D.D. A review of potentially low-cost sorbents for heavy metals. Water Res. 1999, 33, 2469–2479. [Google Scholar] [CrossRef]
  23. Erdem, E.; Karapinar, N.; Donat, R. The removal of heavy metal cations by natural zeolites. J. Colloid Interface Sci. 2004, 280, 309–314. [Google Scholar] [CrossRef] [PubMed]
  24. Baker, H.M.; Massadeh, A.M. Natural Jordanian zeolite: Removal of heavy metal ions from water samples using column and batch methods. Environ. Monit. Assess. 2009, 157, 319–330. [Google Scholar] [CrossRef]
  25. Guo, Z.; Li, Y.; Zhang, S.; Niu, H.; Chen, Z.; Xu, J. Enhanced sorption of radiocobalt from water by Bi(III) modified montmorillonite: A novel adsorbent. J. Hazard. Mater. 2011, 192, 168–175. [Google Scholar] [CrossRef] [PubMed]
  26. Galindo, L.S.G.; Neto, A.F.A. Removal of cadmium(II) and lead(II) ions from aqueous phase on sodic bentonite. Mater. Res. 2013, 16, 515–527. [Google Scholar] [CrossRef] [Green Version]
  27. Ye, X.; Kang, S.; Wang, H.; Li, H.; Zhang, Y.; Wang, G.; Zhao, H. Modified natural diatomite and its enhanced immobilization of lead, copper and cadmium in simulated contaminated soils. J. Hazard. Mater. 2015, 289, 210–218. [Google Scholar] [CrossRef]
  28. Ghodbane, I.; Nouri, L.; Hamdaoui, O.; Chiha, M. Kinetic and equilibrium study for the sorption of cadmium(II) ions from aqueous phase by eucalyptus bark. J. Hazard. Mater. 2008, 152, 148–158. [Google Scholar] [CrossRef]
  29. Bozic, D.; Stankovic, V.; Gorgievski, M.; Bogdanovic, G.; Kovacevic, R. Adsorption of heavy metal ions by sawdust of deciduous trees. J. Hazard. Mater. 2009, 171, 684–692. [Google Scholar] [CrossRef]
  30. Sewwandi, B.G.N.; Vithanage, M.; Wijesekara, S.S.R.M.D.H.R.; Mowjood, M.I.M.; Hamamoto, S.; Kawamoto, K. Adsorption of Cd(II) and Pb(II) onto humic acid–treated coconut (Cocos nucifera) husk. J. Hazard. Toxic Radioact. Waste 2014, 18, 04014001. [Google Scholar] [CrossRef]
  31. Paranavithana, G.N.; Kawamoto, K.; Inoue, Y.; Saito, T.; Vithanage, M.; Kalpage, C.S.; Herath, G.B.B. Adsorption of Cd2+ and Pb2+ onto coconut shell biochar and biochar-mixed soil. Environ. Earth Sci. 2015, 75, 484–494. [Google Scholar] [CrossRef]
  32. Inyang, M.; Gao, B.; Yao, Y.; Xue, Y.; Zimmerman, A.; Mosa, A.; Pullammanappallil, P.; Ok, Y.S.; Cao, X. A review of biochar as a low-cost adsorbent for aqueous heavy metal removal. Crit. Rev. Environ. Sci. Technol. 2016, 46, 406–433. [Google Scholar] [CrossRef]
  33. Chen, X.; Hou, W.H.; Song, G.L.; Wang, Q.H. Adsorption of Cu, Cd, Zn and Pb ions from aqueous solutions by electric arc furnace slag and the effects of pH and grain size. Chem. Biochem. Eng. Q. 2011, 25, 105–114. [Google Scholar]
  34. Agnieszka, G.K.; Baran, P.; Wdowin, M.; Franus, W. Waste dolomite powder as an adsorbent of Cd, Pb(II), and Zn from aqueous solutions. Environ. Earth Sci. 2017, 76, 521. [Google Scholar]
  35. Nguyen, T.C.; Loganathan, P.; Nguyen, T.V.; Kandasamy, J.; Naidu, R.; Vigneswaran, S. Adsorptive removal of five heavy metals from water using blast furnace slag and fly ash. Environ. Sci. Pollut. Res. 2018, 25, 20430–20438. [Google Scholar] [CrossRef] [Green Version]
  36. Gui, C.Y.; Neng, Z.K.; Sheng, Z.Y.; Yue, D.F. Removal of Pb2þ and Cd2þ by adsorption on clay-solidified grouting curtain for waste landfills. J. Cent. South Univ. Technol. 2006, 13, 166–170. [Google Scholar]
  37. Ok, Y.S.; Yang, J.E.; Zhang, Y.S.; Kim, S.J.; Chung, D.Y. Heavy metal adsorption by a formulated zeolite-portland cement mixture. J. Hazard. Mater. 2007, 147, 91–96. [Google Scholar] [CrossRef]
  38. Shaban, E.G.; Isam, M.G.; Abdallah, H.M.G. Cadmium(II) sorption from water samples by powdered marble wastes. Chem. Speciat. Bioavailab. 2015, 20, 249–260. [Google Scholar]
  39. Shaban, E.G.; Abdallah, H.M.G. Lead separation by sorption onto powdered marble waste. Arab. J. Chem. 2014, 7, 277–286. [Google Scholar]
  40. Visa, M.; Chelaru, A.M. Hydrothermally modified fly ash for heavy metals and dyes removal in advanced wastewater treatment. Appl. Surf. Sci. 2014, 303, 14–22. [Google Scholar] [CrossRef]
  41. Yoon, S.; Choi, M.; Hwang, Y.; Bae, S. Upcycling of steel slag for manufacture of Prussian-blue-encapsulated pectin beads and its use for efficient removal of aqueous cesium. J. Clean. Prod. 2021, 319, 128786. [Google Scholar] [CrossRef]
  42. Bibi, S.; Farooqi, A.; Hussain, K.; Haider, N. Evaluation of industrial based adsorbents for simultaneous removal of arsenic and fluoride from drinking water. J. Clean. Prod. 2015, 87, 882–896. [Google Scholar] [CrossRef]
  43. Nguyen, T.C.; Loganathan, P.; Nguyen, T.V.; Vigneswaran, S.; Kandasamy, J.; Naidu, R. Simultaneous adsorption of Cd, Cr, Cu, Pb, and Zn by an iron-coated Australian zeolite in batch and fixed-bed column studies. Chem. Eng. J. 2015, 270, 393–404. [Google Scholar] [CrossRef]
  44. Xin, M.; Yang, S.T.; Tang, H.; Liu, Y.; Wang, H. Competitive adsorption of heavy metal ions on carbon nanotubes and the desorption in simulated biofluids. J. Colloid Interface Sci. 2015, 448, 347–355. [Google Scholar]
  45. Abbas, A.; Al-Amer, A.M.; Laoui, T.; Al-Marri, M.J.; Nasser, M.S.; Khraisheh, M.; Atieh, M.A. Heavy metal removal from aqueous solution by advanced carbon nanotubes: Critical review of adsorption applications. Sep. Purif. Technol. 2016, 157, 141–161. [Google Scholar]
  46. Qiu, Q.; Jiang, X.; Lv, G.; Chen, Z.; Lu, S.; Ni, M.; Yan, J.; Deng, X. Adsorption of heavy metal ions using zeolite materials of municipal solid waste incineration fly ash modified by microwave-assisted hydrothermal treatment. Powder Technol. 2018, 335, 156–163. [Google Scholar] [CrossRef]
  47. Bilal, M.; Ihsanullah, I.; Younas, M.; Shah, M.U.H. Recent advances in applications of low-cost adsorbents for the removal of heavy metals from water: A critical review. Sep. Purif. Technol. 2021, 278, 119510. [Google Scholar] [CrossRef]
  48. Farinella, N.V.; Matos, G.D.; Arruda, M.A.Z. Grape bagasse as a potential biosorbent of metals in effluent treatments. Bioresour. Technol. 2007, 98, 1940–1946. [Google Scholar] [CrossRef]
  49. Pehlivan, E.; Yanık, B.H.; Ahmetli, G.; Pehlivan, M. Equilibrium isotherm studies for the uptake of cadmium and lead ions onto sugar beet pulp. Bioresour. Technol. 2008, 99, 3520–3527. [Google Scholar] [CrossRef]
  50. Kumara, G.M.P.; Saito, T.; Asamoto, S.; Kawamoto, K. Reviews on the applicability of construction and demolition waste as low-cost adsorbents to remove-heavy metals in wastewater. Int. J. Geomate 2018, 14, 44–51. [Google Scholar] [CrossRef]
  51. Yao, Z.T.; Ji, X.S.; Sarker, P.K.; Tang, J.H.; Ge, L.Q.; Xia, M.S.; Xi, Y.Q. Earth-Science Reviews A comprehensive review on the applications of coal fly ash. Earth-Sci. Rev. 2015, 141, 105–121. [Google Scholar] [CrossRef] [Green Version]
  52. Gao, D.; Wang, F.P.; Wang, Y.T.; Zeng, Y.N. Sustainable utilization of steel slag from traditional industry and agriculture to catalysis. Sustainability 2020, 12, 9295. [Google Scholar] [CrossRef]
  53. Reddy, K.R.; Gopakumar, A.; Chetri, J.K. Critical review of applications of iron and steel slags for carbon sequestration and environmental remediation. Rev. Environ. Sci. Biotechnol. 2019, 18, 127–152. [Google Scholar] [CrossRef]
  54. Yusuf, M.; Chuah, L.; Khan, M.A.; Choong, T.S. Adsorption of nickel on electric arc furnace slag: Batch and column studies. Sep. Sci. Technol. 2014, 49, 388–397. [Google Scholar] [CrossRef]
  55. Iucolano, F.; Campanile, A.; Caputo, D.; Liguori, B. Sustainable management of autoclaved aerated concrete wastes in gypsum composites. Sustainability 2021, 13, 3961. [Google Scholar] [CrossRef]
  56. Narayanan, N.; Ramamurthy, K. Structure and properties of aerated concrete: A review. Cem. Concr. Compos. 2000, 22, 321–329. [Google Scholar] [CrossRef]
  57. Hoda, A.; Salarirad, M.M.; Behnamfard, A. Characterization and utilization of clay-based construction and demolition wastes as adsorbents for zinc (II) removal from aqueous solutions: An equilibrium and kinetic study. Environ. Prog. Sust. Energy 2014, 33, 777–789. [Google Scholar]
  58. Taha, A.A.; Wu, Y.; Wang, H.; Li, F. Preparation and application of functionalized cellulose acetate/silica composite nano-fibrous membrane via electrospinning for Cr(VI) ion removal from aqueous solution. J. Environ. Manag. 2012, 112, 10–16. [Google Scholar] [CrossRef]
  59. Chuang, C.L.; Fan, M.; Xu, M.; Brown, R.C.; Sung, S.; Saha, B.; Huang, C.P. Adsorption of arsenic(V) by activated carbon prepared from oat hulls. Chemosp 2005, 61, 478–483. [Google Scholar] [CrossRef] [PubMed]
  60. Monarrez-Cordero, B.E.; Trevizo, A.S.; Carrillo, L.M.B.; Vidaurri, L.G.S.; Yoshida, M.M.; Madrid, P.A. Simultaneous and fast removal of As3+, As5+, Cd2+, Cu2+, Pb2+ and F from water with composite Fe-Ti oxides nanoparticles. J. Alloys Comp. 2018, 757, 150–160. [Google Scholar] [CrossRef]
  61. Kumara, G.M.P.; Kawamoto, K.; Saito, T.; Hamamoto, S.; Asamoto, S. Evaluation of autoclaved aerated concrete (AAC) fines for removal of Cd(II) and Pb(II) from wastewater. J. Environ. Eng. 2019, 145, 04019078. [Google Scholar] [CrossRef]
  62. Gustavo, F.C.; Affonso, C.C.J.; Juan, C.N.M.; David, F.C. Competitive and non-competitive cadmium, copper and lead sorption/desorption on wheat straw affecting sustainability in vineyards. J. Clean. Prod. 2016, 139, 1496–1503. [Google Scholar]
  63. Park, J.H.; Ju, S.C.; Yong, S.O.; Seong, H.K.; Jong, S.H.; Ronald, D.D.; Dong, C.S. Comparison of single and competitive metal adsorption by pepper stem biochar. Arch. Agron. Soil Sci. 2016, 62, 617–632. [Google Scholar] [CrossRef]
  64. Xue, Y.; Hou, H.; Zhu, S. Competitive adsorption of copper(II), cadmium(II), lead(II) and zinc(II) onto basic oxygen furnace slag. J. Hazard. Mater. 2009, 162, 391–401. [Google Scholar] [CrossRef] [PubMed]
  65. Kumara, G.M.P.; Matsuno, A.; Nga, T.T.V.; Giang, N.H.; Kawamoto, K. Simultaneous removal of Pb(II) and Cd(II) from binary and multi-metals solutions using autoclaved aerated concrete and steel slag grains as low-cost adsorbents. In Proceedings of the 17th Sardinia Conference, Forte Village, Santa Margherita Di Pula, Italy, 30 September–4 October 2019. [Google Scholar]
  66. Nippon Slag Association. About Iron and Steel Slag. 2003. Available online: https://www.slg.jp/e/slag/product/doboku.html (accessed on 10 September 2021).
  67. Trong, L.N.; Asamoto, S.; Matsui, K. Sorption isotherm and length change behavior of autoclaved aerated concrete. Cem. Concr. Compos. 2018, 94, 136–144. [Google Scholar] [CrossRef]
  68. Shabalala, A.N.; Stephen, O.E.; Souleymane, D.; Fitsum, S. Pervious concrete reactive barrier for removal of heavy metals from acid mine drainage-column study. J. Hazard. Mater. 2017, 323, 641–653. [Google Scholar] [CrossRef]
  69. Yuanming, S.; Li, B.; Yang, E.H.; Liu, Y.; Ding, T. Feasibility study on utilization of municipal solid waste incineration bottom ash as aerating agent for the production of autoclaved aerated concrete. Cem. Concr. Compos. 2015, 56, 51–58. [Google Scholar]
  70. Marzouk, I.; Hannachi, C.; Dammak, L.; Hamrouni, B. Removal of chromium by adsorption on activated alumina. Desalination Water Treat. 2011, 51, 279–286. [Google Scholar] [CrossRef]
  71. Naiya, T.K.; Bhattacharya, A.K.; Das, S.K. Adsorption of Cd (II) and Pb (II) from aqueous solutions on activated alumina. J. Colloid Interface Sci. 2009, 333, 14–26. [Google Scholar] [CrossRef]
  72. Lopez, F.A.; Martin, M.I.; Perez, C.; Lopez-Delgado, A.; Alguacil, F.J. Removal of copper ions from aqueous solutions by a steel-making by-product. Water Res. 2003, 37, 3883–3890. [Google Scholar] [CrossRef]
  73. Organization for Economic Co-operation and Development Publications (OECD). Guidelines for the Testing of Chemicals. 2000. Available online: https://www.oecd.org/chemicalsafety/testing/oecdguidelinesforthetestingofchemicals.htm (accessed on 15 September 2021).
  74. Metcalf & Eddy, Inc.; Tchobanoglous, G.; Burton, F.L.; Stensel, H.D. Chap. 2 Constituents in Wastewater. In Wastewater Engineering: Treatment and Reuse, 4th ed.; McGraw-Hill Series in Civil and Environmental Engineering; McGraw-Hill Science Engineering: Boston, MA, USA, 2002. [Google Scholar]
  75. Shafiq, M.; Alazba, A.A.; Amin, M.T. Kinetic and isotherm studies of Ni2+ and Pb2+ adsorption from synthetic wastewater using eucalyptus camdulensis—derived biochar. Sustainability 2021, 13, 3785. [Google Scholar] [CrossRef]
  76. Saadi, R.; Saadi, Z.; Fazaeli, R. Monolayer and multilayer adsorption isotherm models for sorption from aqueous media. Korean J. Chem. Eng. 2015, 32, 787–799. [Google Scholar] [CrossRef]
  77. Martemianov, D.; Xie, B.B.; Yurmazova, T.; Khaskelberg, M.; Wang, F.; Wei, C.H.; Preis, S. Cellular concrete-supported cost-effective adsorbents for aqueous arsenic and heavy metals abatement. J. Environ. Chem. Eng. 2017, 5, 3930–3941. [Google Scholar] [CrossRef]
  78. Damrongsiri, S. Feasibility of using demolition waste as an alternative heavy metal immobilising agent. J. Environ. Manag. 2017, 192, 197–202. [Google Scholar] [CrossRef]
  79. Kumara, G.M.P.; Kawamoto, K. Applicability of crushed clay brick and municipal solid waste slag as low-cost adsorbents to refine high concentrate Cd (II) and Pb (II) contaminated wastewater. Int. J. Geomate 2019, 17, 133–142. [Google Scholar] [CrossRef]
  80. Bourliva, A.; Michailidis, K.; Sikalidis, C.; Filippidis, A.; Betsiou, M. Adsorption of Cd (II), Cu (II), Ni (II) and Pb (II) onto natural bentonite: Study in mono- and multimetal systems. Environ Earth Sci. 2015, 73, 5435–5445. [Google Scholar] [CrossRef] [Green Version]
  81. Seyedeh, S.H.G.; Landi, A.; Khademi, H.; Hojati, S. Removal of Cd2+ and Pb2+ ions from aqueous solutions using Iranian natural zeolite and sepiolite. J. Environ. Stud. 2014, 40, 43–45. [Google Scholar]
  82. He, S.; Li, Y.; Weng, L.; Wang, J.; He, J.; Liu, Y.; Zhang, K.; Wu, Q.; Zhang, Y.; Zhang, Z. Competitive adsorption of Cd2+, Pb2+ and Ni2+ onto Fe3+-modified argillaceous limestone: Influence of pH, ionic strength and natural organic matters. Sci. Total Environ. 2018, 637–638, 69–78. [Google Scholar] [CrossRef] [PubMed]
  83. Youness, A.; Olguín, M.T.; Abatal, M.; Ali, B.; Mendez, S.E.D.; Santiago, A.A. Comparison of the divalent heavy metals (Pb, Cu and Cd) adsorption behavior by montmorillonite-KSF and their calcium- and sodium-forms. Superlattices Microstruct. 2019, 127, 165–175. [Google Scholar]
  84. Yang, D.; Liu, Y.; Liu, S.; Li, Z.; Tan, X.; Huang, X.; Zeng, G.; Zhou, Y.; Zheng, B.; Cai, X. Competitive removal of Cd(II) and Pb(II) by biochars produced from water hyacinths: Performance and mechanism. RSC Adv. 2016, 6, 5223–5232. [Google Scholar]
  85. Bulgariu, D.; Bulgariu, L. Equilibrium and kinetics studies of heavy metal ions biosorption on green algae waste biomass. Bioresour. Technol. 2012, 103, 489–493. [Google Scholar] [CrossRef] [PubMed]
  86. Arslanoglu, H.; Altundogan, H.S.; Tumen, F. Heavy metals binding properties of esterified lemon. J. Hazard. Mater. 2009, 164, 1406–1413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Duan, J.; Su, B. Removal characteristics of Cd(II) from acidic aqueous solution by modified steel-making slag. Chem. Eng. J. 2014, 246, 160–167. [Google Scholar] [CrossRef]
  88. Dimitrova, S.V. Metal sorption on blast-furnace slag. Water Res. 1996, 30, 228–232. [Google Scholar] [CrossRef]
  89. Dimitrova, S.V.; Mehandgiev, D.R. Lead removal from aqueous solutions by granulated blast-furnace slag. Water Res. 1998, 32, 3289–3292. [Google Scholar] [CrossRef]
  90. Dimitrova, S.V.; Mehandgiev, D.R. Interaction of Blast-furnace slag with heavy metal ions in water solutions. Water Res. 2000, 34, 1957–1961. [Google Scholar] [CrossRef]
  91. Takeno, N. Atlas of Eh-pH Diagrams; Intercomparison of Thermo Dynamic Database. 2005. Available online: https://www.gsj.jp/researches/openfile/openfile2005/openfile0419.html (accessed on 3 September 2021).
Figure 1. Measured adsorption isotherms and fitted Langmuir and Freundlich models for Cd2+ (a,b) and Pb2+ (c,d) onto three grain sizes of SS.
Figure 1. Measured adsorption isotherms and fitted Langmuir and Freundlich models for Cd2+ (a,b) and Pb2+ (c,d) onto three grain sizes of SS.
Sustainability 13 10321 g001
Figure 2. Relationship between released Ca2+ amount and adsorbed metal amount of SS grain (0.105–2 mm). Measured from adsorption isotherms in single-metal solution of Cd2+ (0 ≤ Ci ≤ 2000 mg/L) at L/S = 60 and Pb2+ (0 ≤ Ci ≤ 1500 mg/L) at L/S = 10.
Figure 2. Relationship between released Ca2+ amount and adsorbed metal amount of SS grain (0.105–2 mm). Measured from adsorption isotherms in single-metal solution of Cd2+ (0 ≤ Ci ≤ 2000 mg/L) at L/S = 60 and Pb2+ (0 ≤ Ci ≤ 1500 mg/L) at L/S = 10.
Sustainability 13 10321 g002
Figure 3. Metal removal %, R, in binary-metal solution (Cd2+ and Pb2+) for tested adsorbents with different mixing proportions (grain size = 0.105–2 mm). (ae) x-axis is Cd:Pb ratio in binary-metal solution; (fj) x-axis is Pb:Cd ratio in binary-metal solution. Measured pH values after the adsorption (pHe) ranged 8–10 for AAC and 10–12 for SS and mixed adsorbents of SS and AAC.
Figure 3. Metal removal %, R, in binary-metal solution (Cd2+ and Pb2+) for tested adsorbents with different mixing proportions (grain size = 0.105–2 mm). (ae) x-axis is Cd:Pb ratio in binary-metal solution; (fj) x-axis is Pb:Cd ratio in binary-metal solution. Measured pH values after the adsorption (pHe) ranged 8–10 for AAC and 10–12 for SS and mixed adsorbents of SS and AAC.
Sustainability 13 10321 g003
Figure 4. Schematics of simultaneous removal of Cd2+ and Pb2+ by mixing grains of SS and AAC in binary-metal solution (alkaline condition).
Figure 4. Schematics of simultaneous removal of Cd2+ and Pb2+ by mixing grains of SS and AAC in binary-metal solution (alkaline condition).
Sustainability 13 10321 g004
Figure 5. Metal removal %, R, in multi-metal solution for (a) SS, (b) SS+AAC [4:1], (c) SS+AAC [1:1], (d) SS+AAC [1:4], and (e) AAC in different L/S conditions. Measured pH values after the adsorption (pHe) ranged from 8–10 for AAC and from 10–12 for SS and mixed adsorbents of SS and AAC.
Figure 5. Metal removal %, R, in multi-metal solution for (a) SS, (b) SS+AAC [4:1], (c) SS+AAC [1:1], (d) SS+AAC [1:4], and (e) AAC in different L/S conditions. Measured pH values after the adsorption (pHe) ranged from 8–10 for AAC and from 10–12 for SS and mixed adsorbents of SS and AAC.
Sustainability 13 10321 g005
Table 1. Basic physical and chemical properties of tested adsorbents.
Table 1. Basic physical and chemical properties of tested adsorbents.
AdsorbentParticle Size
(mm)
pHEC
(mS/cm)
LOI
(%)
wAD
(%)
GsSSA
(m2/g)
Reference
DI1M KCl
<0.10512.913.16.114.52.53.058.0
SS0.105–212.412.66.94.81.94.9This study
2–4.7513.113.15.04.81.92.6
<0.10510.08.71.710.34.4 23.6
AAC0.105–210.08.31.810.83.92.4923.6[61]
2–4.759.98.91.49.84.0 21.9
Table 2. Chemical composition of tested adsorbents.
Table 2. Chemical composition of tested adsorbents.
AdsorbentComposition (wt %)Reference
SiO2CaOAl2O3Fe2O3K2OMgOOthers
SS16.841.73.222.50.23.512.1This study
AAC48.633.82.81.90.50.412.0[61]
Table 3. Summary of batch adsorption experiments and testing conditions.
Table 3. Summary of batch adsorption experiments and testing conditions.
AdsorbentTypes of ExperimentLiquid-to-Solid Ratio (L/S)Initial Metal Concentration, (Ci = mg/L)
SS1. Adsorption isotherm in single-metal solution (Cd2+, Pb2+) 60 (Cd2+), 10 (Pb2+)0–5000 (Cd2+), 0–1500 (Pb2+)
SS
SS+AAC [4:1]
SS+AAC [1:1]
SS+AAC [1:4]
AAC
2. Removal % in binary-metal solution (Cd2+ + Pb2+)601000
3. Removal % in multi-metal solution (Cd2+ + Pb2+ + Cu2+ + Ni2+ + Zn2+)601000
4. Effect of L/S ratio on removal % in multi-metal solution 5, 10, 60, 100, 2501000
Note: Deionized water (DI) with natural pH was used as a background solution for all experiments.
Table 4. Fitted Langmuir and Freundlich model parameters for Cd2+ and Pb2+ adsorption onto tested adsorbents.
Table 4. Fitted Langmuir and Freundlich model parameters for Cd2+ and Pb2+ adsorption onto tested adsorbents.
AdsorbentMetalParticle Size (mm)LangmuirFreundlichReference
Qe (mg/g)b (L/mg)r2Kf (mg/g)1/nr2
SSCd2+<0.10531311050.270.98This study
0.105–22440.180.991140.170.89
2–4.751300.090.9992.00.040.77
Pb2+<0.10517.50.020.970.570.670.97
0.105–28.60.010.910.150.700.96
2–4.758.20.010.930.130.720.94
AAC <0.10516.51.720.999.890.110.91[61]
Cd2+0.105–216.50.910.999.920.100.93
2–4.7515.20.690.998.820.110.89
<0.10525811370.200.97
Pb2+0.105–225711210.240.98
2–4.752500.050.9974.30.220.98
Note: 1:Qm values were estimated by the measured maximum adsorption amount. Langmuir and Freundlich parameters for SS were determined fitting the adsorption isotherm of Cd2+ at 0 ≤ Ci ≤ 5000 mg/L and the adsorption isotherm of Pb2+ at 0 ≤ Ci ≤ 1500 mg/L. Langmuir and Freundlich parameters for AAC were determined fitting the adsorption isotherms of Cd2+ and Pb2+ at 0 ≤ Ci ≤ 2000 mg/L.
Table 6. Selectivity sequence of metal removal, R (%), for tested adsorbents with different liquid-to-solid ratio (L/S).
Table 6. Selectivity sequence of metal removal, R (%), for tested adsorbents with different liquid-to-solid ratio (L/S).
AdsorbentL/SSelectivity SequenceR (%)
Cd2+Pb2+Cu2+Ni2+Zn2+
SS5Cd2+ ≈ Cu2+ ≈ Ni2+ > Zn2+ > Pb2+10086.110010099.8
10Cd2+ ≈ Cu2+ ≈ Ni2+ > Zn2+ > Pb2+10086.910010099.7
60Pb2+ > Cu2+ > Cd2+ > Ni2+ ≈ Zn2+55.299.563.420.920.6
100Pb2+ > Cu2+ > Cd2+ > Zn2+ > Ni2+51.776.263.80.98.0
250Pb2+ > Cd2+ > Cu2+ > Zn2+ > Ni2+45.962.031.70.06.1
SS+AAC [4:1]5Cd2+Pb2+ ≈ Cu2+ ≈ Ni2+ ≈ Zn2+ 100100100100100
10Cd2+ ≈ Cu2+ ≈ Ni2+ > Pb2+ > Zn2+ 10096.410010044.5
60Cu2+ > Pb2+ > Cd2+ > Zn2+ > Ni2+ 46.187.899.317.923.2
100Pb2+ > Cu2+ > Cd2+ > Zn2+ > Ni2+38.360.142.90.06.3
250Cd2+ > Pb2+ > Cu2+ > Zn2+ > Ni2+39.137.128.40.25.3
SS+AAC [1:1]5Cd2+Pb2+ ≈ Cu2+ ≈ Ni2+ ≈ Zn2+100100100100100
10Cd2+> ≈ Pb2+ ≈ Cu2+ ≈ Zn2+ > Ni2+10010010095.299.9
60Pb2+ ≈ Cu2+ > Cd2+ > Zn2+ > Ni2+28.699.899.811.520.3
100Pb2+ > Cu2+ > Cd2+ > Zn2+ > Ni2+26.859.939.30.25.4
250Pb2+ > Cd2+ > Cu2+ > Zn2+ > Ni2+26.445.522.30.04.3
SS+AAC [1:4]5Ni2+ > Cu2+ > Zn2+Pb2+ > Cd2+99.499.599.910099.8
10Cu2+ > Pb2+ > Zn2+ > Cd2+ > Ni2+47.299.210023.971.4
60Cu2+ > Pb2+ > Cd2+ > Zn2+ > Ni2+21.997.698.50.06.5
100Pb2+ > Cu2+ > Cd2+ > Zn2+ > Ni2+18.469.043.40.01.3
250Pb2+ > Cu2+Cd2+ > Zn2+ > Ni2+17.839.618.00.00.1
AAC5Pb2+ ≈ Cu2+ > Zn2+ > Cd2+ > Ni2+41.310010044.771.1
10Cu2+ > Pb2+ > Ni2+ > Cd2+ > Zn2+32.199.410032.920.6
60Pb2+ > Cu2+ > Cd2+ ≈ Ni2+ > Zn2+19.410060.117.95.3
100Pb2+ > Cu2+ > Cd2+ > Zn2+ > Ni2+15.060.441.02.94.2
250Pb2+ > Cu2+ > Cd2+ > Zn2+ > Ni2+14.030.725.20.02.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumara, G.M.P.; Kawamoto, K. Steel Slag and Autoclaved Aerated Concrete Grains as Low-Cost Adsorbents to Remove Cd2+ and Pb2+ in Wastewater: Effects of Mixing Proportions of Grains and Liquid-to-Solid Ratio. Sustainability 2021, 13, 10321. https://doi.org/10.3390/su131810321

AMA Style

Kumara GMP, Kawamoto K. Steel Slag and Autoclaved Aerated Concrete Grains as Low-Cost Adsorbents to Remove Cd2+ and Pb2+ in Wastewater: Effects of Mixing Proportions of Grains and Liquid-to-Solid Ratio. Sustainability. 2021; 13(18):10321. https://doi.org/10.3390/su131810321

Chicago/Turabian Style

Kumara, Gajanayake Mudalige Pradeep, and Ken Kawamoto. 2021. "Steel Slag and Autoclaved Aerated Concrete Grains as Low-Cost Adsorbents to Remove Cd2+ and Pb2+ in Wastewater: Effects of Mixing Proportions of Grains and Liquid-to-Solid Ratio" Sustainability 13, no. 18: 10321. https://doi.org/10.3390/su131810321

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop